You are on page 1of 45

J. Fluid Mech. (1999), vol. 398, pp. 109153.

Printed in the United Kingdom


c 1999 Cambridge University Press
109
Simulation of boundary layer transition induced
by periodically passing wakes
By XI AOHUA WU
1
, ROBERT G. J ACOBS
1
,
J ULI AN C. R. HUNT
2
AND PAUL A. DURBI N
1
1
Center for Integrated Turbulence Simulations, Flow Physics and Computation Division,
Department of Mechanical Engineering, Stanford University, Building 500, Stanford,
CA 94305-3030, USA
e-mail: wu@transition.stanford.edu; rjacobs@vk.stanford.edu; durbin@vk.stanford.edu
2
Department of Applied Mathematics and Theoretical Physics,
University of Cambridge, Cambridge CB3 9EW, UK
e-mail: jrh2@hermes.cam.ac.uk
(Received 9 September 1998 and in revised form 1 June 1999)
The interaction between an initially laminar boundary layer developing spatially on
a at plate and wakes traversing the inlet periodically has been simulated numer-
ically. The three-dimensional, time-dependent NavierStokes equations were solved
with 5.24 10
7
grid points using a message passing interface on a scalable parallel
computer. The ow bears a close resemblance to the transitional boundary layer on
turbomachinery blades and was designed following, in outline, the experiments by Liu
& Rodi (1991). The momentum thickness Reynolds number evolves from Re

= 80
to 1120. Mean and second-order statistics downstream of Re

= 800 are of canonical


at-plate turbulent boundary layers and are in good agreement with Spalart (1988).
In many important aspects the mechanism leading to the inception of turbulence is
in agreement with previous fundamental studies on boundary layer bypass transition,
as summarized in Alfredsson & Matsubara (1996). Inlet wake disturbances inside
the boundary layer evolve rapidly into longitudinal pus during an initial receptivity
phase. In the absence of strong forcing from free-stream vortices, these structures ex-
hibit streamwise elongation with gradual decay in amplitude. Selective intensication
of the pus occurs when certain types of turbulent eddies from the free-stream wake
interact with the boundary layer ow through a localized instability. Breakdown of
the pus into young turbulent spots is preceded by a wavy motion in the velocity
eld in the outer part of the boundary layer.
Properties and streamwise evolution of the turbulent spots following breakdown,
as well as the process of completion of transition to turbulence, are in agreement with
previous engineering turbomachinery ow studies. The overall geometrical character-
istics of the matured turbulent spot are in good agreement with those observed in the
experiments of Zhong et al. (1998). When breakdown occurs in the outer layer, where
local convection speed is large, as in the present case, the spots broaden downstream,
having the vague appearance of an arrowhead pointing upstream.
The ow has also been studied statistically. Phase-averaged velocity elds and skin-
friction coecients in the transitional region show similar features to previous cascade
experiments. Selected results from additional thought experiments and simulations are
also presented to illustrate the eects of streamwise pressure gradient and free-stream
turbulence.
110 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
Tangential
Axial
y
x
Rotor
Wake
Stator
U
rotor
U
o
u
t
U
r
e
f
U
wake
Figure 1. Sketch of rotorstator wake interaction; U
rotor
: rotor velocity in the stator reference
frame; U
out
: rotor exit ow velocity in the rotor reference frame; U
ref
: stator inow velocity in the
stator reference frame.
1. Introduction
Upstream and downstream blade interactions in the passages of multi-stage axial
turbomachines result in a complex and inherently unsteady ow eld. Specically, the
boundary layer over a blade surface is subjected to a substantial degree of unsteadiness
that stems from impinging wakes of the upstream stator or rotor (gure 1). The full
turbomachinery conguration includes other disturbances, but wake impingement is
the dominant eect at subsonic Mach numbers (Korakianitis 1993; Hodson 1998).
The pronounced eect of upstream wakes arises primarily because large regions of
laminar and transitional ows exist on the suction surface of embedded stages. The
impinging vortical, turbulent wake markedly alters the path to transition.
There are a number of mechanisms for the way in which eddies and organized
disturbances interact with an adjacent boundary layer that is initially laminar. These
have been identied in recent research, but their combined eect in any given ow
is dicult to understand and to predict. The rst question is to dene which of the
entrainment or external mechanisms are relevant: do perturbations in the boundary
layer emanate from an upstream edge or inlet and grow within the boundary layer, or
are internal disturbances induced directly by external disturbances that move above
the layer? This is the general question of receptivity paths for bypass transition
(Goldstein & Wundrow 1998; Leib, Wundrow & Goldstein 1999; Jacobs & Durbin
1998). In the case of incident turbulent wakes, large external downdrafts or gusts
Boundary layer transition induced by passing wakes 111
may be deected by the vorticity of the layer through the sheltering mechanism of
Hunt & Durbin (1999), or they may penetrate the boundary layer to produce locally
amplifying turbulent spots.
The second question concerns how the uctuations are transformed within the
boundary layer and how the layer itself is changed. Instabilities may be triggered at
subcritical Reynolds numbers (Corral & Jimenez 1994) through the action of nite-
amplitude disturbances. Only very low level forcing produces transition via Tollmein
Schlichting waves. Moderate or high level forcing leads to transition via formation
of localized turbulent spots without TollmeinSchlichting precursors (Mayle 1991).
Once induced, these disturbances grow within the boundary layer, although their
development may be inuenced by modulation of the boundary layer by the free-
stream distortion.
The specic receptivity path and internal growth mechanics depend on the particular
ow conguration (Hunt & Durbin 1999). However, general questions can be asked
within the scope of transition induced by localized, convected external disturbances
that either enter the ow domain abruptly or are rapidly distorted at a leading edge.
There are several mechanisms operative in passing wake-induced bypass transition.
The mean wakes distort boundary layer proles and the wakes are distorted by the
wall: does this signicantly alter the receptivity and transition processes? Fluctuations
are created as the turbulent wake enters the ow domain: are these the origin of tur-
bulent spots, or do their long-wavelength components just modulate the downstream
transition mechanism? The convected wakes carry free-stream turbulence over the
boundary layer: are these free-stream eddies the primary source of transitional spots?
The present simulations address these questions.
This paper describes a spatially developing, three-dimensional, time-accurate DNS
of boundary layer transition induced by periodically passing wakes (gure 2). The
incident wakes are generated as self-similar free shear ows, but the manner of their
introduction into the ow domain requires an inevitable distortion near the wall.
However, that distortion is well dened and reproducible.
1.1. The connection with some fundamental work on bypass transition
There is now a sizable body of literature from fundamental studies of boundary layer
bypass transition due to moderate-amplitude free-stream turbulence. It was realized
by us, only retrospectively, that the problem at hand shares important features with
some of the previous experimental, theoretical and numerical studies. These common
characteristics are related to the physical mechanisms leading to the inception of
turbulent spots, their growth through the transition region, and the manner in which
they maintain the downstream turbulent region.
Experiments by Alfredsson & Matsubara (1996) in a laminar boundary layer
subjected to 1.5% and 6% free-stream turbulence showed that during the initial
receptivity and evolution phase, free-stream turbulence induces longitudinal streaks
with a fairly periodic, spanwise regularity inside the boundary layer. These structures
grow downstream both in length and amplitude. Breakdown to turbulent spots was
observed to occur in the regions where smoke visualization exhibited intensive streaks.
The breakdown of streaks often occurs after a wavy motion of the streaks, although
spots occur locally and abruptly, not via amplication of the wavy motion to the point
of breakdown. The turbulent spots grow in number and size downstream, until the
boundary layer becomes fully turbulent. Similar streaky structures inside laminar and
transitional boundary layers were observed by Grek, Kozlov & Ramazanov (1985)
and termed pus; streaks were also found in transitional channel ow (Klingmann
112 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
x
u
wake
x
wake
y
wake
v
wake

U
ref
y
L
U
cyl
(a)
x
U
ref
=1
y
L =1
U
cyl
= 0.7 U
ref
(b)
0
Figure 2. (a) Layout in the experiments of Liu & Rodi (1991); (b) layout in the present numerical
simulation; the computational domain is dened as 0.1 6 x/L 6 3.5, 0 6 y/L 6 0.8, 0 6 z/L 6 0.2.
1992; Henningson, Lundbladh & Johansson 1993) and in ows experiencing oblique
transition (Berlin, Lundbladh & Henningson 1994).
Previous research on bypass transition induced by free-stream turbulence left open
to interpretation the issue of whether intensication and ultimate destruction of the
streaky structures arises from boundary layer internal dynamics, or from forcing by
free-stream eddies. Such ambiguity is primarily due to experimental diculties in
following the details of the generation and growth of disturbances because of their
randomness in space and time.
Figure 3(ac) illustrates three scenarios observed in the present investigation (see
, 4). In the most interesting case of those drawn in gure 3, inlet wake disturbances
rapidly evolve into pus similar to those found in Westin et al. (1994) and Alfredsson
& Matsubara (1996) and turbulent spots appear downstream. More usually, the pus
decay as in gure 3(b), at least below the critical Reynolds number Re

= 200.
Turbulent eddies inside the passing free-stream wake impinge on the boundary layer
and sometimes interact with its outer part, such as to subject the ow to a rapidly
growing instability (, 4). This involves an intensication of the near-wall streaky
structures, and eventually the breakdown into young turbulent spots. Figures 3(d)
and 3(e) depict two other scenarios in which the ow is subjected to very strong
or extremely weak disturbances from the passing wake. Observations concerning the
Boundary layer transition induced by passing wakes 113
(a)
Puff
Forcing
Breakdown
(young spot)
Turbulent
spot
Turbulent
strip
Further decay Decay
(b)
(c)
Forcing
Young spot Decay
(d)
Turbulent patch
(e)
Elongated structure with spanwise modulation
z
x
Figure 3. Several possibilities for the downstream propagation of certain types of inlet distur-
bances. The sketches represent u
/
in an (x, z)-plane very close to the wall: (a) intermediate-strength
disturbance and strong forcing; (b) intermediate-strength disturbance and weak forcing; (c) inter-
mediate-strength disturbance with downstream strong forcing; (d) strong disturbance; (e) very weak
disturbance.
last two instances have been made in the present study through additional numerical
simulations and thought experiments (see , 7).
1.2. The connection with some engineering turbomachinery research
Experiments on wake-induced periodic unsteady transition in turbomachinery blade
rows were reported by Dring et al. (1982), Dong & Cumpsty (1990), Addison &
Hodson (1990), Mayle & Dullenkopf (1991), and Halstead et al. (1997) (see also
114 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
the reviews by Mayle 1991 and Walker 1993). Halstead et al. (1996) reported mea-
surements in compressors and low-pressure turbines. Their experiments showed that
transition in unsteady, turbomachine boundary layers develops along two dierent,
but coupled, paths. These consist of a wake-induced strip under the convecting wake
trajectory, and a path between wakes that is caused by other disturbances. Along
both paths the boundary layer goes from laminar to transitional to turbulent, with
large regions of laminar and transitional ow. The switch from the non-wake path
to the wake-induced strip was found to occur in a small fraction of a blade passing
period. Halstead et al. noted that assumptions of predominantly turbulent boundary
layers on multi-stage turbomachine blading are incorrect.
Although turbine and compressor experiments have indicated that the eects of
wake passing can be substantial and have provided useful guidelines for further
research, the technical complexities involved in obtaining detailed quantitative data
from rotating turbine/compressor stages make it dicult to isolate physical mecha-
nisms. Realizing such complexities, a number of investigators have considered simpler
geometries. In the simplest of these (Pfeil, Herbst & Schroder 1983; Liu & Rodi 1991;
Orth 1993; Zhong et al. 1998) the unsteady blade row interaction was simulated by
sweeping a row of wake-generating cylinders past a at plate (gure 2). Liu & Rodi
(1991) obtained time- and phase-averaged mean and uctuating streamwise velocity
proles for four dierent wake passing frequencies. In their experiments, the Reynolds
number was fairly low so that the boundary layer remained laminar over the full plate
length when no disturbing wakes were present. They found that the wake-produced
turbulent strips grew together and caused the boundary layer to become fully turbu-
lent. The streamwise location of the merger moved upstream with increasing wake
passing frequency.
Using experimental data gathered from a similar ow conguration, Orth (1993)
concluded that in turbomachinery ows periodically disturbed by passing wakes,
the disturbance enters the boundary layer very early on, and convects within it
before leading to transition. Periodic uctuation in the velocity prole, as opposed
to stochastic uctuation, does not have a major inuence on the transition. This is
consistent with the present study. Orth (1993) also suggested that the location where
transition takes place is only dependent on inlet turbulence intensity: the passing
wake exerts no eect on the process. Our study shows that this needs qualication.
Inception of turbulent spots in wake-induced transitional ows is intimately linked
with turbulent eddies of the travelling free-stream wake. One additional pleasant
connection of the present study with turbomachinery research concerns the recent
liquid crystal visualization experiments of Zhong et al. (1998) and Kittichaikarn et
al. (1999). This work was communicated to us by Prof. Hodson. Our turbulent spots,
as well as their embryo precursors, resemble those observed by Zhong et al. (1998)
and Kittichaikarn et al. (1999) to a remarkable degree (see , 3 and , 4).
2. Mathematical and numerical considerations
2.1. Problem denition
Consider the evolution of an incompressible ow over a smooth at plate with
upstream wakes passing periodically (gure 2b). The origin of the coordinate system
is at the leading edge of the plate. The wakes are assumed to be generated by
imaginary circular cylinders positioned in the plane x = L and moving in the y-
direction at U
cyl
, which can be either positive or negative, corresponding to inlet wakes
Boundary layer transition induced by passing wakes 115
traversing away from, or towards, the at plate. The velocity of the ow upstream
of the cylinder is U
ref
. The cylinders are equally spaced so that they cut through
the y = 0 plane at a specied passing period . The characteristic velocity scale is
U
ref
, the characteristic length scale is L, the Reynolds number is then Re = U
ref
L/
where is the kinematic viscosity of the uid. Throughout this study Re = 1.5 10
5
,
as in Liu & Rodi (1991). The mean ow properties of the wake are determined by
free-stream velocity, cylinder velocity and cylinder diameter.
The computational domain for the DNS is dened as 0.1 6 x/L 6 3.5, 0 6 y/L 6
0.8, and 0 6 z/L 6 0.2. The inlet momentum thickness Reynolds number is Re

= 80
in all the simulations. Unless otherwise noted, all velocities are normalized by the
reference velocity U
ref
and all lengths by the characteristic length scale L.
2.2. Governing equations and notation
Mass and momentum conservation is enforced for ow over the at plate by solving
the full time-dependent, mass-conservation and NavierStokes equations in Cartesian
coordinates,
div u = 0, (1)
u
t
+ div (uu) =
1

grad p + div
_
1
Re
_
grad u + (grad u)

_
, (2)
where u is the velocity vector with Cartesian components (u, v, w) or u
i
, i = 1, 2, 3.
Superscript denotes transpose. The equations are in non-dimensional form.
In this paper, time-averaging is represented by . Averaging at a particular phase,
t
m
n

= m+ n

, is denoted by (), where m is any integer and 0 6 n

6 1 is the
fraction of the wake passing period. For example, the phase-averaged mean velocity
components are evaluated as
(u
i
)(t
n

) =
1
M
M

m=1
u
i
(t
m
n

), (3)
where M is the total number of periods within which phase averaging is per-
formed. Averaging over the homogeneous spanwise z-direction is implied in both
time-averaging and phase-averaging. Time-averaged and phase-averaged mean veloc-
ities are related via u
i
= (u
i
). Thus the instantaneous velocity can be decomposed
as
u
i
= (u
i
)(t
n

) + u
/
i
(t
n

) = u
i
+ u
i
(t
n

) + u
/
i
(t
n

), (4)
where u
i
(t
n

) = (u
i
)(t
n

) u
i
is the periodic velocity uctuation with respect to
the time-averaged mean, and u
/
i
(t
n

) is the true stochastic turbulence uctuation.


Consequently, the time-averaged Reynolds stresses (u
/
i
u
/
j
) can be calculated as
(u
/
i
u
/
j
) =
_
1
0
_
u
i
(u
i
)(t
n

)
_
u
j
(u
j
)(t
n

)
_
dn

. (5)
2.3. Inow and other boundary conditions
For all the simulations described in this paper, the height of the computational
domain 0.8L is approximately 200 at the inlet x = 0.1L, 20 in the middle of the
plate x = 1.75L, and 11 at the exit x = 3.5L. (x) is the 99% boundary layer
thickness. The width of the computational domain 0.2L is equivalent to 40 at the
inlet, 5 in the middle of the plate and 3 at the exit.
116 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
Depending upon the passing frequency 1/, at any given instant either one single
wake or multiple wakes can be found on the inow plane. A wake will be found
on the computational inow plane if its centreline is located within the range from
y = b cos to y = 0.8L+b cos , where b is wake half-width and = tan
1
U
cyl
/U
ref
is the wake inclination angle (gure 2). The total number of such wakes at any time
is
.= CEILING
_
0.8L + 2b cos
U
cyl

_
, (6)
where the function CEILING returns the smallest integer greater than or equal to its
real argument.
Consider U
cyl
< 0. On the inow plane at the beginning of each wake passing
period (t
m
n

= 0), let the origin of the uppermost wake-coordinate system (x


wake, 1
= 0,
y
wake, 1
= 0) start to move downwards from y = 0.8L + b cos . This indicates that
at t
m
n

= 0 the wake starts to enter the top boundary of the inow plane. At any
subsequent time 0 < t
m
n

< the origin of the wake-coordinate system will be


at y
centreline
= 0.8L + b cos + t
m
n

U
cyl
. Each point on the inow plane then has an
eective coordinate with respect to the origin of the wake-coordinate system, which
is y
e, wake
(y) = (y y
centreline
)/cos .
For U
cyl
> 0, the origin of the wake-coordinate system starts to move upwards from
y = b cos . This indicates that at t
m
n

= 0 the wake starts to enter the lower boundary


of the inow plane. At any subsequent time the origin of its wake-coordinate system
will be at y
centreline
= b cos + t
m
n

U
cyl
. Each point on the inow plane then has an
eective y
wake
-coordinate with respect to the origin of the wake-coordinate system,
which is y
e, wake
(y) = (y y
centreline
)/cos .
At the inow station of the computational domain x = 0.1L, the velocity compo-
nents were prescribed as
u = u
blasius
+ u
blasius
_
cos
.

q=1
u
e, wake, q
sin
.

q=1
v
e, wake, q
_
,
v = v
blasius
+ u
blasius
_
sin
.

q=1
u
e, wake, q
+ cos
.

q=1
v
e, wake, q
_
,
w =
.

q=1
w
e, wake, q
,
_

_
(7)
where subscript blasius denotes the steady Blasius prole. The subscript e, wake, q
represents the eective instantaneous velocity components in the qth wake-coordinate
system. The velocities (u
e, wake, q
, v
e, wake, q
, w
e, wake, q
) are the same as the wake velocities
(u
wake
, v
wake
, w
wake
) illustrated in gure 2 if y
e, wake, q
6 b, and zero otherwise. The
wake velocities were multiplied by u
blasius
so they would satisfy no slip where the wake
intersects the plate. As the wake enters the computational domain, a more realistic
prole develops rapidly. Distortion of the velocities at the inlet results in a localized
streamwise pressure gradient. The eect of such a pressure gradient on transition and
boundary layer development will be studied in , 7 through numerical experiments.
In general, the mean velocity prole in wakes at a large distance from a solid
body is independent of the shape of the body, except for a scale factor (Schlichting
1979). Liu & Rodi (1991) also pointed out that near the leading edge of the plate
the cylinder wake is self-similar, and has also lost its shedding characteristics. Raj &
Lakshminarayna (1973) demonstrated that the wake behind an airfoil trailing edge is
Boundary layer transition induced by passing wakes 117
similar to that of a two-dimensional cylinder wake. Invoking self-similarity allows the
wake introduced at the inlet to be obtained more simply than by actually computing
a real cylinder wake.
The turbulent wake velocities (u
wake
, v
wake
, w
wake
) appearing in (7) were generated
from a separate precomputation of a temporally decaying, self-similar plane wake,
following the work of Moser, Rogers & Ewing (1998) and Ghosal & Rogers (1997) .
In a temporally decaying wake, the ow is statistically homogeneous in the streamwise
and spanwise directions, and inhomogeneous in the cross-stream direction. The initial
conditions for the temporally decaying plane wake simulation were generated from a
turbulent channel ow simulation at Reynolds number 3300 based on the centreline
mean velocity and channel half-height. The procedure involved taking two realizations
of half-channel ow and fusing them together. Physically this corresponds to a
situation in which two half-channel ows exist on either side of a rigid plate and
the plate is instantaneously removed. This simulation was performed on a grid size
of (65, 128, 128) using an LES code (Wu & Squires 1997). The grid sizes used by
Ghosal & Rogers (1997) and Moser et al. (1998) were (65, 48, 16) and (512, 195, 128),
respectively.
Mean ow and turbulence statistics of the simulated plane wake are presented in
gure 4. All velocities in the gure are normalized by the maximum mean velocity
decit u
wake, max
; lengths are normalized by the wake half-width b. Following Ghosal
& Rogers (1997) , the half-width b is dened as the distance between the two points at
which the mean velocity decit is 50%u
wake, max
. This is slightly larger than the distance
between the wake centreline and the rst point with eectively zero mean velocity
decit. Figure 4(a) shows that mean velocity proles obtained at the three dierent
instants (indicated in the caption by their descending maximum velocity decits)
collapse. These mean proles are also in excellent agreement with experimental
measurements and the data correlation of Schlichting (1979), i.e. u
wake
/u
wake, max
=
[1 (y
wake
/1.1338b)
1.5
]
2
. This demonstrates that the simulated wake has reached a
self-similar state and that the mean ow has lost its memory of the initial condition.
At these and subsequent instants, the product of the wake width and maximum decit
remains constant.
Figure 4(b, c) show the r.m.s. turbulence intensities for the decaying plane wake
at the same three instants. Turbulent shear stress proles are given in gure 4(d).
The turbulence intensities obtained by Moser et al. (1998) and Ghosal & Rogers
(1997) have two distinct features: double peaks occur in the streamwise component,
and the wall-normal component is slightly higher than the spanwise component.
In addition, the results of Ghosal & Rogers show that unlike the mean ow and
the anti-symmetrical Reynolds shear stress proles, the turbulence intensities are
not self-similar as time proceeds. It is evident from gure 4(bd) that the present
precomputation reproduced all these essential features. Figure 4(e) presents proles
of the rate of turbulence kinetic energy dissipation. Spanwise energy spectra of the
turbulence kinetic energy are given in gure 4(f) for completeness. Note that the
results in gure 4 are presented in the wake-coordinate system (see gure 2). The
uctuating wake velocities obtained from the precomputation are rescaled by the
wake maximum decit u
wake, max
and half-width b before they are applied to (7).
Using the experimental correlation of Schlichting (1979) for far wakes, at x/L = 0.1
u
wake, max
= 0.14U
ref
and b 0.1L for the present ow conditions.
At the top surface of the computational domain the following boundary conditions
were applied: v = v
blasius
, u/y = v/x, and w/y = v/z. This is articial, but
given the substantial distance between the top surface and the wall, the eect of the
118 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(e)
0.20
0.16
0.12
0.08
0.04
0
2 1 0 1 2
y/b

b
/
u
3w
a
k
e
,

m
a
x
( f )
10
10
1 10 100 1000
k
z
b
(c)
0.8
0.6
0.4
0.2
0
2 1 0 1 2
y/b
(d)
0.10
0.06
0.02
0.02
0.06
0.1
2 1 0 1 2
y/b
(a)
0.1
0.1
0.3
0.5
0.7
1.1
2 1 0 1 2
y/b
(b)
0.8
0.6
0.4
0.2
0
2 1 0 1 2
y/b
10
8
10
6
10
4
10
2
10
0
E
/
b
u
2w
a
k
e
,

m
a
x
0.9
u
w
a
k
e
/
u
w
a
k
e
,

m
a
x
u

2
1
/
2
w
a
k
e
,

w

2
1
/
2
w
a
k
e
/
u
w
a
k
e
,

m
a
x
u
w
a
k
e

v
w
a
k
e
/
u
2w
a
k
e
,

m
a
x
v

2
1
/
2
w
a
k
e
/
u
w
a
k
e
,

m
a
x
Figure 4. Characteristics of the simulated temporally decaying plane wake for the generation of
inow turbulence: , u
wake, max
= 0.12; , u
wake, max
= 0.10; , u
wake, max
= 0.08; , plane
cylinder wake of Schlichting (1979); , E k
5/3
z law. (a) Mean velocity; (b) streamwise and
spanwise uctuations; (c) wall-normal uctuations; (d) turbulent shear stress; (e) viscous dissipation
rate of turbulence kinetic energy; (f) spanwise spectrum of turbulence kinetic energy.
top boundary condition on boundary layer development should be extremely small.
At the exit of the computational domain, convective boundary conditions were used.
Mass ux at the inow plane was made constant in time by rescaling the velocities
obtained from (7), and corrections to the velocities at the exit plane are also made
to ensure global mass conservation. Periodic boundary conditions were applied in the
homogeneous, spanwise z-direction; u = 0 was applied on the wall.
Boundary layer transition induced by passing wakes 119
(c)
2.0
1.5
1.0
0.5
0 0.02 0.04 0.06 0.08
y

1
/
2
,
+
(d)
2.0
1.5
1.0
0.5
0 0.02 0.04 0.06 0.08
y
(a)
0.8
0.6
0.4
0.2
0
10
1
(b)
3
2
1
0 0.02 0.04 0.06 0.08

1
/
2
,
+

1
/
2
,
+
10
2
10
3
10
4
10
5
1.0
u
Figure 5. Resolution check: symbols, baseline case with x
+
x =3
= 24 (x
+
x =1
= 18.3) and z
+
x =3
= 11
(z
+
x =1
= 8.4):
e
, during transition x = 1.0; , after transition x = 3.0; , spanwise resolution
rened by 50%; , streamwise resolution coarsened by 50%.
2.4. Numerical method
The numerical scheme for the DNS is a parallelized version (by Charles D. Pierce at
Stanford) of the method used by Akselvoll & Moin (1996) and Pierce & Moin (1998).
Convection and diusion terms that involve only derivatives in the wall-normal
direction are treated implicitly, whereas all other terms are treated explicitly. All
spatial derivatives are approximated with a second-order central dierence scheme.
A third-order RungeKutta scheme (Spalart, Moser & Rogers 1991) is used for
terms treated explicitly and a second-order CrankNicolson scheme is used for terms
treated implicitly. The fractional step method is used to remove the implicit pressure
dependence in the momentum equations. Further details can be found in Akselvoll
& Moin (1996). For parallel computation the computational domain is decomposed
in two directions whereas a third direction is complete. When solving the Poisson
equation using fast transforms, a transpose is necessary to switch the un-decomposed
direction. Scalable parallelization is achieved using message passing interface (MPI)
libraries.
2.5. Computational details and resolution check
The governing equations are solved on a rectangular staggered grid. The grid spacings
are uniform in the streamwise and spanwise directions.
Simulation results to be presented in the following sections were obtained on a
(1024, 400, 128) grid in the streamwise, wall-normal and spanwise directions, respec-
120 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c) t/= 32.9
0.6
0.4
0 0.5 1.0 1.5
x
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2 wake 1
(b) t/= 32.7
0.6
0.4
0 0.5 1.0 1.5
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2 wake 1
(a) t/= 32.5
0.6
0.4
0 0.5 1.0 1.5
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2
wake 1
Figure 6. Contours of u over one (x, y)-plane.
tively. The total of 52.4 million grid points is one of the largest that has ever been
reported: compare to 11.0 million, 6.2 million and 17.3 million used by Spalart (1988),
Yang, Spalart & Ferziger (1992) and Rai & Moin (1993), respectively.
In terms of viscous wall units based on the time-averaged local friction velocity
after transition, x
+
x =3
= 24 and z
+
x =3
= 11. When measured using a friction velocity
during transition, x
+
x =1
= 18.3 and z
+
x =1
= 8.4. At the exit, there are 16 points
distributed along the wall-normal direction below y
+
= 9, and a total of 191 points
below y = . The resolution used in Spalart (1988) was x
+
20_1, z
+
6.7_0.34,
with 10 points within 9 wall units.
In order to check the adequacy of the streamwise and spanwise grid resolution,
two complete additional simulations were performed. It was dicult to use a grid size
larger than (1024, 400, 128) due to memory constraints of the computer. Therefore, in
the rst additional simulation the spanwise dimension of the computational domain
was reduced from 0.2L to 0.15L, which is equivalent to 30 at the inlet, 3.94 in the
middle of the plate and 2.18 at the exit. Even such a reduced spanwise dimension
is still suciently wide and we therefore assume most of the dierences, if any,
between the two sets of results are due to the spanwise grid resolution change from
z
+
x =1
= 8.4 to z
+
x =1
= 6.3. In the second additional simulation, the number of grid
points in the streamwise direction was reduced by 50% from 1024 to 768. Except for
these changes, all the other parameters were kept the same as the baseline simulation.
Results from the resolution check are presented in gure 5. Figure 5(a) compares the
three sets of mean velocity proles at two streamwise stations: the rst, at x = 1.0,
is in the transitional region; and the second, at x = 3.0, is in the turbulent region.
Boundary layer transition induced by passing wakes 121
(c) t/= 32.9
0.6
0.4
0 0.5 1.0 1.5
x
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2 wake 1
(b) t/= 32.7
0.6
0.4
0 0.5 1.0 1.5
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2 wake 1
(a) t/= 32.5
0.6
0.4
0 0.5 1.0 1.5
0.2
y
2.0 2.5 3.0 3.5
wake 3
wake 2
wake 1
Figure 7. Contours of v over one (x, y)-plane. In this and subsequent similar gures, negative
values are contoured by solid lines; positive values are contoured by dashed lines.
The curves show good numerical resolution. Figures 5(b), 5(c) and 5(d) compare
the streamwise, wall-normal and spanwise r.m.s. turbulent intensities, respectively.
Dierences among the three simulations are small. Additional resolution checks can
be found in , 5. In addition to these resolution checks, we also build condence on our
simulation through extensive comparison, presented in , 5, with well-accepted DNS
and experimental data. Comparison with the resolution used in previous channel ow
turbulent spot simulations (e.g. Henningson & Kim 1991) gives further condence that
the present resolution is adequate. Years of turbulence simulation research at Stanford
University has shown that, despite the slow convergence rate with grid renement,
second-order central dierencing has several attractive features. It is energy conserving
and does not carry inherent numerical diusion, as do many high-order upwind biased
schemes.
The time step was xed to be dt = 10
3
=0.00167L/U
ref
, which is equivalent to
0.59 /u
2
, x =3
. Initial velocities were set to the laminar Blasius prole. The ow was
then allowed to evolve for 20 wake passing periods (20 000 dt), and statistics were
then collected for another 20 wake passing periods. Phase averaging was performed
by dividing each pass period into 50 equal subdivisions. The computation was carried
out on the scalable parallel Cray T3E at the Pittsburgh Supercomputing Center, using
up to 512 processors.
3. Visualization of a matured turbulent spot
At the beginning of each period t
m
n

= 0, the passing wake starts to enter the


computational inow plane at the top boundary y = 0.8L. Since the cylinder travels
122 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(a)
0.10
0.05
0.625 0.750
x
0.00
z
0.875 1.000 1.125
(b)
Flow
Figure 8. (a) Contours of u
/
at t/ = 32.5 over the (x, z)-plane y = 7.38 10
4
(y
+
= 5.4 at
x = 1.75) negative u
/
represented by dashed lines; (b) visualization from the experiments of Zhong
et al. (1998).
at 0.7U
ref
, it takes 1.143L/U
ref
or 0.684 before the wake reaches the at plate.
After entering the computational domain, the wake is advected at the reference
velocity U
ref
in the free stream and interacts with the boundary layer in the near-wall
region. An important grid resolution requirement is that it be adequate to ensure that
no vorticity is spuriously left in the free stream after the wake has passed.
An overall view of the wake evolution as it is advected along the plate and the
laminar to turbulent transition can be seen in gures 6 and 7. Figure 6 presents
contours of instantaneous streamwise velocity over one random (x, y)-plane at three
consecutive instants: t = 32.5, 32.7 and 32.9. It is seen from the gure that in
the free stream the wake angle remains the same at all the three instants, and that
there is no residual velocity gradient left by the passing wake. The small eects of the
upper boundary show that the computational domain is suciently high compared to
the boundary layer thickness. Between the wakes the near-wall velocity contours from
the inlet to x 1.0 are straight, indicating that the ows in these moving regions are
predominantly laminar. Beyond x = 1.75 the contours are chaotic in the near-wall
region at all the instants, and there is also apparent thickening in the boundary layer.
This indicates that laminar-to-turbulent transition has been nearly completed.
Figure 7 presents contours of instantaneous wall-normal velocity v at the same
locations and instants as in gure 6. The wall-normal velocity inside the wake is
signicantly larger than that in the free stream of a normal laminar or turbulent
boundary layer because of the wake angle . The gure shows that downstream
Boundary layer transition induced by passing wakes 123
0.2
0.1
1.75 2.00 2.25
x
0
z
2.50 2.75
Figure 9. Contours of u
/
at t/ = 32.5 over the (x, z)-plane of y = 7.38 10
4
(y
+
= 5.4 at
x = 1.75) showing turbulent streaks in the turbulent region negative u
/
represented by darker
contours.
Wall
Speaker
y
x
U
b
Breakdown (source)
U
U
a
A
y
x
A
U
b
U
a
Breakdown (source)
z
x
Wall
y
x
U
b
Breakdown
U
U
a
A
y
x
A
U
b
U
a
Breakdown
z
x
Wall
Wall
Figure 10. One possible reconciliation of the spot arrowhead direction found in
the present study with that in previous boundary layer studies.
0.05
0.03
0.7 0.8 0.9
x
y
1.0 1.1
0.01
1.2
Figure 11. Side-view of the turbulent spot at t/= 32.5 over the (x, y)-plane of z = 0.03: ,
0.20U
ref
< v
/
< 0.01U
ref
; - - - -, 0.01U
ref
< v
/
< 0.20U
ref
; increment 0.005U
ref
.
of x 2.0 the instantaneous wall-normal velocity is chaotic near the wall, with
substantial magnitude all the time, as would be found inside a turbulent boundary
layer. It is interesting to note that, upstream of x 1.75L, an isolated spot near
the wall containing large-amplitude and chaotic vertical velocities is developing on
124 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c) t/= 32.9
0.2
0.1
0 0.5 1.0 1.5
x
z
2.0 2.5 3.0 3.5
(b) t/= 32.7
0.2
0.1
0 0.5 1.0 1.5
z
2.0 2.5 3.0 3.5
(a) t/= 32.5
0.2
0.1
0 0.5 1.0 1.5
z
2.0 2.5 3.0 3.5
Figure 12 (ac). For caption see facing page.
the upstream side of the free-stream wake. It can be seen from the contours that as
the spot propagates downstream, its streamwise dimension lengthens. The evolution
of this particular turbulent spot, as well as its connection to the overall unsteady
boundary layer transition, will be examined in the remainder of this subsection.
Figure 8(a) provides a close-up plan view of the turbulent spot indicated by
gure 7(a). Shown in the gure are contours of u
/
at t = 32.5 over the (x, z)-
plane of y = 7.3810
4
. The spot has an arrowhead shape pointing upstream, with
streamwise elongation. Because u
/
represents turbulence uctuations with respect to
the phase-averaged mean, the contours are predominantly positive inside the spot. If
u
/
is computed with respect to a conditional mean, averaged only inside the spot, there
will be both positive and negative uctuations. This can also be inferred from the
high-speed and low-speed streaks inside the spot in gure 8(a). Figure 8(b) presents
a ow visualization picture obtained by Zhong et al. (1998) in an experimental
conguration similar to ours (gure 2b). Good agreement between the DNS spot and
the experimental visualization is evident. Figure 9 shows contours of u
/
at the same
time and (x, z)-plane as in gure 8(a), but within a further downstream, fully turbulent
region. The most distinctive feature found in the gure is the existence of low- and
high-speed streaks. These streaks and their associated streamwise vortices have been
recognized as a signature of fully developed near-wall turbulence (Hamilton, Kim &
Boundary layer transition induced by passing wakes 125
( f ) t/= 33.6
0.2
0.1
0 0.5 1.0 1.5
x
z
2.0 2.5 3.0 3.5
(e) t/= 33.5
0.2
0.1
0 0.5 1.0 1.5
z
2.0 2.5 3.0 3.5
(d) t/= 33.4
0.2
0.1
0 0.5 1.0 1.5
z
2.0 2.5 3.0 3.5
Figure 12. Visualization of spot growth and transition to turbulence using v
/
over the (x, z)-
plane of y = 7.38 10
4
(y
+
= 5.4 at x = 1.75); contours represent 0.005U
ref
6 v
/
6 0.2U
ref
with
increment 0.005U
ref
.
Walee 1995). The commonly accepted characteristic wavelength of such streaks is
about 100 wall units. The entire spanwise dimension in gure 9 is about 1400 wall
units. Such turbulent streaks should not to be confused with the pus existing in the
laminar region prior to the occurrence of turbulent spots, as drawn in gure 3.
Our turbulent spot and that of Zhong et al. (1998) have an arrowhead pointing
upstream, in the reverse direction to that reported in many previous boundary layer
studies (Henningson, Spalart & Kim 1987; Jahanmiri, Prabhu & Narasimha 1996).
This discrepancy is too glaring to be left uncommented on.
Figure 10 depicts a simple rationale for the dierence. The breakdown uid parcels
contain the source of turbulence, which is spread to form a spot as the parcels
themselves are convected downstream. In many previous boundary layer turbulent
spot studies breakdown takes place near the wall, where the local convection velocity
is small. At the position, A, located higher above the wall than the source, the
turbulent uid parcels travel downstream relative to the source. At any given instant
uid parcels at smaller x
A
have a longer time to spread laterally, thus forming an
arrowhead pointing downstream.
126 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
Breakdown provoked by free-stream turbulence occurs in the outer part of the
layer where the local convection velocity is large (see , 4 for further discussion). In
the reference frame of observer A, nearer to the wall than the breakdown location,
the highly turbulent breakdown uid parcels are convected downstream. At any given
instant, uid parcels at larger x
A
have had longer to spread laterally, thus forming an
arrowhead pointing upstream. Alternatively, in the frame of reference of the source,
the boundary layer ow is increasingly in the x-direction as the wall is approached.
Thus the spot is sheared towards the upstream direction near the wall: the vertical
section in gure 11 illustrates this structure.
Evolution of a turbulent spot and the process of unsteady periodic transition is
portrayed in gure 12. Presented in this gure are visualizations of the instantaneous
wall-normal uctuation v
/
over the (x, z)-plane, y = 7.38 10
4
, which is equivalent
to y
+
x =1.75
= 5.4. A total of six consecutive realizations are given from t = 32.5
to t = 33.6, covering more than a full wake passing period. In the gure, the
background is used to represent negligible uctuations with 0.005 6 v
/
6 0.005.
Other contours represent stronger uctuations 0.005 < v
/
< 0.2 with an increment
of 0.005. Note that the spanwise dimension in gure 12 has been magnied by a
factor of four in order to show the full streamwise dimension. The spot appearing in
gure 12(a) is the same as that presented in gure 8(a), except that the contours are
now drawn using v
/
. Because of the enlarged z-dimension, the spot dimensions are
distorted. Their real physical shape is as in gure 8(a).
In gure 12(b) a small patch of large turbulent uctuations exists near the upper
boundary (z 0.2, x 1.0). This is part of the wing tip of the turbulent spot near
the lower boundary, extended via the periodic boundary condition in the spanwise
direction. From t = 32.5 to 32.7 the arrowhead of the turbulent spot broadens,
but is still pointing upstream. In the meantime the ank of the spot becomes sharper,
making a well-dened angle with respect to the ow direction. As the wing tip
expands, more uctuations stronger than the background level are seen to appear
near z = 0.2 as well. Along the z = 0 boundary, the downstream turbulent region
retreats from x 1.65 at 32.5 to x 1.75 at 32.7.
At t = 32.7 the next wake starts to touch the at plate. Large turbulence
uctuations advected into the near-wall region of the computational domain by the
wake can be seen in gure 12(c) near x = 0.25. These uctuations decay rapidly
so that in gure 12(d) they have entirely disappeared. This provides evidence that
turbulent spots are not produced immediately where impact occurs.
From t = 32.7 to 32.9, the shape of the turbulent spot transforms from a
well-dened wedge to a two-dimensional strip. This process occurs through a rapid
increase in the angle which the ank makes with the streamwise direction. Eventually
the arrowhead shape disappears, resulting from a tendency towards equalization of
the dimensions of the leading and trailing edges of the spot.
Formation of a full two-dimensional strip, and catching up the downstream tur-
bulent region by the strip, are shown in gure 12(d). The two-dimensional strip does
not extend over the whole spanwise dimension until t = 33.4, nearly 0.9 passing
period after the appearance of the spot in gure 12(a). Relative to the given pass-
ing frequency, the evolution from isolated turbulent spots to a full two-dimensional
strip is rather gradual. Prior to being caught by the turbulent strip, the continuously
turbulent region had moved to x 2.4L; through the act of being caught, the contin-
uously turbulent region jumps back to 1.6L, restoring its position of gure 12(a). The
transition cycle is begun anew by the emergence of two small turbulent spots near
x 0.9L in gure 12(e). Figure 12(f) shows the state of these two young turbulent
spots after another 0.1 passing period.
Boundary layer transition induced by passing wakes 127
(a)
0.10
0.05
0.0 0.25
x
z
0.5 0.75
(b)
Flow
x
z
0.15
0.20
Figure 13. Pu prior to the emergence of turbulent spot: (a) u
/
at t/= 33.2 over the (x, z)-plane
of y = 7.38 10
4
(y
+
= 5.4 at x = 1.75). (b) Visualization from the experiments of Kittichaikarn
et al. (1999).
4. The search for the origin of a young turbulent spot
We have argued in , 1 and , 3 that inlet disturbances develop into streaky structures
(pus) inside the boundary layer. Selective amplication of the pus in the transitional
region occurs when certain types of free-stream wake eddies interact with boundary
layer ow. Breakdown starts near the boundary layer edge. We are now ready to
provide supportive evidence for this argument.
Figure 13(a) demonstrates the existence of streaky structures in the transitional
region. Shown in the gure are u
/
contours at t = 33.2 over the same (x, z)-
plane as in gure 12. This particular instant is 0.3 prior to the emergence of
young spots in gure 12(e). The elongated positive and negative u
/
contours resem-
ble the pu sketched in gure 3, and are essentially the same as those observed
by Alfredsson & Matsubara (1996) in a boundary layer under continuous free-
stream turbulence. Figure 13(b) presents one interesting ow visualization picture
obtained by Kittichaikarn et al. (1999) that compares favourably with our simulation.
The three streaky structures are precursors of turbulent spots. Additional visualiza-
tions show the occurrence of three well-dened turbulent spots some time following
gure 13(b).
Figure 14 follows the three pus of gure 13(a) through their earlier history. It starts
with their inception, near the inlet, and follows them to their ultimate breakdown
128 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c) t/= 33.1
0.20
0.15
0 0.25
x
z
0.50 0.75
0.10
0.05
(b) t/= 32.9
0.20
0.15
0 0.25
z
0.50 0.75
0.10
0.05
(a) t/= 32.8
0.20
0.15
0 0.25
z
0.50 0.75
0.10
0.05
Figure 14 (ac). For caption see facing page.
near x = 1.0. That is the point of the emergence of young spots in gure 12(e). In
this gure the contour plots of v
/
have the threshold level reduced by a factor of
ten compared to gure 12 (from 0.005 to 0.0005). Near-wall inlet wake disturbances
shown in gure 14(a) rapidly evolve into three patches in gure 14(b). Each of these
three patches is made of elongated structures with upward and downward motions,
suggesting an association with streamwise vortices. Figure 14(c) shows that the top
pu (near z = 0.175) and lower pu (near z = 0.025) decay from 32.9 to 33.1,
whereas the middle pu (near z = 0.1) intensies. Attenuation or amplication of
the near-wall pu is dependent upon the type of forcing from free-stream eddies.
Figure 15 illustrates such forcing. Figure 15(a) shows u
/
in an (x, y)-plane cut through
the centre of the top pu, at an instant midway between gure 14(b) and gure 14(c).
The time-averaged boundary layer thickness in this region is about 0.006. Positive
u
/
are found behind the wake inside the boundary layer (0.25 < x < 0.4), implying
the absence of inectional instability. Figure 15(c) shows u
/
in an (x, y)-plane cut
through the centre of the lower pu. Again, positive u
/
is evident. The lack of forcing
through inectional instability in gures 15(a, c) corresponds to the decay of the top
Boundary layer transition induced by passing wakes 129
( f ) t/= 33.5
0.20
0.15
0.50 0.75
x
z
1.00 1.25
0.10
0.05
(e) t/= 33.4
0.20
0.15
0.50 0.75
z
1.00 1.25
0.10
0.05
(d) t/= 33.3
0.20
0.15
0.50 0.75
z
1.00 1.25
0.10
0.05
Figure 14. Visualization of the evolution of pus and breakdown to young turbulent spots using
v
/
over the (x, z)-plane of y = 7.38 10
4
(y
+
= 5.4 at x = 1.75); 0.0005U
ref
6 v
/
6 0.008U
ref
with
increment 0.0005.
and lower pu shown in gure 14(c). Figure 15(b) presents contours of u
/
in an
(x, y)-plane cut through the centre of the middle pu. Evident from the gure are
the negative u
/
contours in the region where free-stream wake eddies interact with
boundary layer ow. This corresponds to the amplication of the middle pu seen in
gure 14(c).
Figure 14(df) shows variations of the three pus before breakdown. The trend
toward attenuation of the top pu is reversed. This is because the free-stream forcing
shown in gure 16(a) is of the opposite sign to that in gure 15(a). u
/
is strongly
negative. Strong negative u
/
is also seen in gure 16(b). Under such forcing, the top
and middle pu break up at t = 33.5. In contrast, gure 16(c) shows there is very
weak forcing in the lower pu. This corresponds to its monotonic decay as seen in
gure 14(bf). Figure 17 visualizes the evolution of the three pus by way of contours
of streamwise uctuations.
Figure 18 shows the breakdown using contours of u
/
. Again, the three (x, y)-planes
z = 0.175, 0.11 and 0.025 cut through the centres of the three pus. Due to the
negative u
/
, the boundary layers in gure 18(a) and gure 18(b) have inectional
130 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c) z = 0.025, t/= 33.0
0.25 0.5
x
y
0.75 1.0
0.10
0.08
0.06
0.04
0.02
(b) z = 0.11, t/= 33.0
0.25 0.5
y
0.75 1.0
0.10
0.08
0.06
0.04
0.02
(a) z = 0.175, t/= 33.0
0.25 0.5
y
0.75 1.0
0.10
0.08
0.06
0.04
0.02
Figure 15. Visualization of the forcing by free-stream eddies through localized instability using u
/
over three (x, y)-planes at t/=33.0; 0.02U
ref
6 u
/
6 0.34U
ref
with increment 0.04.
proles. The wavy structures in the velocity eld are the signature of their instability.
Figure 18(b) suggests that breakdown occurs rst in the outer part of the layer. When
breakdown takes place, the negative streamwise uctuations are replaced by positive
uctuations (compare gure 18(b) and gure 19(b) at x = 1.0). A turbulent spot
is made of predominantly positive u
/
(see gure 8a), thus high local skin friction.
Near the downstream end of the turbulent spot, positive u
/
of the spot collides with
negative u
/
associated with the forcing (see gure 19b at x = 1.05). Continuity thus
results in a high positive wall-normal uctuation v
/
in this region. The strong upward
motion near the downstream end of the spot results in an overhang, which is clearly
visible in gure 7(c).
One nal observation to be made is that gure 18(b) shows more conclusively
that breakdown starts from the outer layer than gure 18(a). Our reasoning in , 3
concerning the arrowhead direction of the spot associated with the middle pu is that
it should have a more well-dened arrowhead pointing upstream than the top one.
This is indeed the case. See the two young spots in gure 12(f) and note again that
the spanwise dimension has been enlarged by a factor of four.
Boundary layer transition induced by passing wakes 131
(c) z = 0.025, t/= 33.4
0.75 1.00
x
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
(b) z = 0.11, t/= 33.4
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
(a) z = 0.175, t/= 33.4
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
Figure 16. Visualization of the forcing by free-stream eddies through localized instability
using u
/
over three (x, y)-planes at t/=33.4; 0.02U
ref
6 u
/
6 0.34U
ref
with increment 0.04.
5. Time-averaged boundary layer properties
Time-averaged boundary layer integral parameters are presented in gure 20. In
addition to the laminar Blasius solution, two sets of DNS results are shown: the
simulation with z
+
x =3
= 11 and an additional simulation with z
+
x =3
= 8.25. The
laminar-ow momentum thickness Reynolds number Re

is 80 at the inow station


(x = 0.1). This is dictated by the assumption that the Blasius boundary layer is
initiated from the leading edge with the prescribed length Reynolds number. At
the exit of the computational domain (x = 3.5) the turbulent boundary layer has
Re

= 1120. It is seen from the gure that upstream of x 0.75 the simulations
follow the Blasius solution quite nicely with only very small deviations. The minor
dierences are due to the impact of the wake on the at plate. Onset of transition
starts at about x 0.7 and by x 2.0 the shape factor has dropped from 2.59
to 1.45. Further downstream the shape factor remains nearly the same, decreasing
only slightly from 1.45 to 1.42 at the exit. Coles (1956) correlation shows that in a
turbulent boundary layer the shape factor drops from 1.48 at Re

= 600 to 1.44 at
Re

= 1100.
Quantitative time-averaged skin-friction data are not available in most experiments
132 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c) t/= 33.1
0 0.25
x
z
0.50 0.75
0.20
0.15
0.10
0.05
(b) t/= 32.9
0 0.25
z
0.50 0.75
0.20
0.15
0.10
0.05
(a) t/= 32.8
0 0.25
z
0.50 0.75
0.20
0.15
0.10
0.05
Figure 17 (ac). For caption see facing page.
on wake-induced transition. Nevertheless, wall shear stress information is crucial since
it provides an important velocity scale for boundary layer theory as well as a necessary
quantity for engineering drag estimation. Figure 21 shows C
f
. Similar to the previous
gure, the Blasius solution and the results from the two simulations are presented.
Within 0.1 6 x 6 0.5, proles of the computed skin-friction coecient follow the
Blasius solution with only a minor over-shoot near the inlet because of the impact of
the wake on the at-plate. C
f
starts to rise beyond x 0.7 and attains a maximum
at approximately 2.15. The streamwise location of the maximum skin friction may
be used as a convenient, well-dened, indicator for the completion of the transition
process. As will be shown next, time-averaged mean streamwise velocity and Reynolds
shear stresses attain their corresponding fully turbulent proles at approximately the
same streamwise station. At the exit, x = 3.5, the computed skin-friction coecient C
f
is 0.00479, 10% higher than that given by Coles correlation. Free-stream turbulence
uctuations, such as those carried by the passing wake in the present case, tend to
increase skin-friction (Hancock & Bradshaw 1989).
Time-averaged streamwise velocities at seven streamwise stations are shown in
gure 22. Figure 22(a) plots u
+
in inner coordinates. The three proles upstream of x =
Boundary layer transition induced by passing wakes 133
( f ) t/= 33.5
0.50 0.75
x
z
1.00 1.25
0.20
0.15
0.10
0.05
(e) t/= 33.4
0.50 0.75
z
1.00 1.25
0.20
0.15
0.10
0.05
(d) t/= 33.3
0.50 0.75
z
1.00 1.25
0.20
0.15
0.10
0.05
Figure 17. Visualization of the evolution of pus using u
/
over the (x, z)-plane of
y = 7.38 10
4
(y
+
= 5.4 at x = 1.75); 0.02U
ref
6 u
/
6 0.14U
ref
with increment 0.02.
1.5 display large deviations from the standard logarithmic prole u
+
= 2.44 ln y
+
+5.0,
though the degree of deviation decreases along the streamwise direction. At x = 2.0
the prole of u
+
still does not possess a well-dened logarithmic slope, indicating that
on average transition is not complete at this station. Further downstream, the three
proles at x = 2.5, 3.0 and the exit (3.5) nearly collapse within 0 6 y
+
6 300. In the
viscous region they follow the law of the wall u
+
= y
+
. In the logarithmic region the
slopes of these proles are in excellent agreement with that of the log law, i.e. 1/
with = 0.41. The intercept of the proles is lower than the standard value 5.0 by
approximately 0.8. This might be attributed to the higher time-averaged skin-friction
value discussed in gure 21. The logarithmic velocity proles produced by the present
simulation are clearly dened. Interestingly, the wake component (Coles 1956) in
the outer part of the boundary layer is also well-dened even though Hancock &
Bradshaw (1989) showed this tends to be suppressed when the intensity of free-stream
turbulence exceeds the friction velocity u

. Our numerical experiments show that the


absence or existence of Coles wake component depends upon wake orientation and
passing frequency (see Wu & Durbin 1999b for further discussion).
134 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
( c) z = 0.025
0.75 1.00
x
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
( b) z = 0.11
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
( a) z = 0.175
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
Figure 18. Visualization of the forcing by free-stream eddies through localized instability using u
/
over three (x, y)-planes at t/=33.5; 0.02U
ref
6 u
/
6 0.34U
ref
with increment 0.04.
Figure 22(b) plots u/U
ref
in outer coordinates. Also shown in the gure are the
experimental measurements of Webster, DeGra & Eaton (1996) on a at plate at
Re

= 1500. In the transitional region, the proles of u/U


ref
become fuller with the
increase of streamwise coordinate. A noticeable characteristic of these proles is that
velocities in the inner and outer regions of the boundary layer approach the fully
turbulent experimental data dierently. Close to the wall u increases monotonically
with x. However, there is an over-shoot of u in the outer region of the boundary
layer. The degree of over-shoot is the largest at x = 1.5, where the boundary layer is
in the midst of its transitional state (see gure 27), and where there are large wall-
normal velocity gradients in the region connecting the inner and outer parts of the
boundary layer. The over-shoot decreases further downstream as the ow becomes
fully turbulent. At the exit x = 3.5 (Re

= 1120) the computed mean velocity


prole u/U
ref
is in excellent agreement with the experimental data of Webster et al.
Figure 22(c) plots u/U
ref
with respect to y to show the absolute change. Away from
the wall the variation with streamwise distance is monotonic, corresponding to the
growth of the boundary layer shown in gure 20. The inner part drops from x = 0.5
to 1.0 rst, before u starts to increase. This is entirely consistent with the skin friction
variation shown in gure 21.
Boundary layer transition induced by passing wakes 135
( c) z = 0.025
0.75 1.00
x
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
( b) z = 0.11
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
( a) z = 0.175
0.75 1.00
y
1.25 1.50
0.10
0.08
0.06
0.04
0.02
Figure 19. Visualization of the forcing by free-stream eddies through localized instability using u
/
over three (x, y)-planes at t/=33.6; 0.02U
ref
6 u
/
6 0.34U
ref
with increment 0.04.
Time-averaged turbulence intensities and Reynolds shear stress at six streamwise
stations are presented in gure 23, together with the DNS data of Spalart (1988)
obtained from a turbulent at-plate boundary layer at Re

= 1410. Figure 23(a)


shows that even at the early stage of transition x = 0.5 there exist relatively large
streamwise uctuations in the outer part of the boundary layer, but the levels of
wall-normal intensity and Reynolds shear stress are very low. This is consistent with
previous studies discussed by Alfredsson & Matsubara where streamwise intensity
was found to have a maximum in the centre of the boundary layer. At x = 1.0 and
1.5 the proles of (u
/
2
)
1/2,+
exhibit large over-shoots in the outer region compared to
the fully turbulent prole. Signicantly the peaks are also located away from the wall,
e.g. 0.15 for x = 1.0. Unlike the three turbulence intensities, Reynolds shear stress
increases from x = 0.5 to 1.5 almost monotonically throughout the boundary layer
and asymptotes to that of Spalart (1988). Note, however, that all the proles shown
in gure 23 are normalized by the local friction velocity, which masks the absolute
changes between dierent streamwise stations (see gure 25). Figure 23(e) shows that
at x = 2.5, after the end of transition, proles of turbulence intensities and Reynolds
136 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
0.5 1.5
x
I
n
t
e
g
r
a
l

p
a
r
a
m
e
t
e
r
s
2.5
12
10
4
2
0 1.0 2.0 3.0 3.5
8
6
Figure 20. Time-averaged mean boundary layer integral parameters: , Blasius solution without
wake; , simulation with z
+
x =3
= 11 (z
+
x =1
= 8.4); symbols, simulation with z
+
x =3
= 8.25
(z
+
x =1
= 6.3):

, 10
2
; , 10
2

; .,

/; , 10
2
Re

.
0.5 1.5
x
C
f
2.5
0.006
0.005
0.002
0.001
0 1.0 2.0 3.0 3.5
0.004
0.003
Figure 21. Time-averaged mean skin-friction coecient: , Blasius solution without wake;
, simulation with z
+
x =3
= 11 (z
+
x =1
= 8.4); , simulation with z
+
x =3
= 8.25 (z
+
x =1
= 6.3).
shear stress are in very good agreement with Spalarts DNS. As expected, the free-
stream intensities are higher than Spalart (1988) because of the passing wake. From
x = 2.5 on downstream, changes in the proles are minimal. At the exit (gure 23f)
the maximum value of (u
/
2
)
1/2,+
and its location are in excellent agreement with
Spalart (1988). The prole also develops a shoulder near 0.15 which is commonly
found in low Reynolds number boundary layer ows.
Boundary layer transition induced by passing wakes 137
0.04
y
0.06
1.0
0.4
0.2
0 0.02
0.8
0.6
(c)
u
1.0
y/
1.2
1.0
0.4
0.2
0
0.6
0.8
0.6
(b)
u
10
y
+
100
30
10
0
1
20
(a)
1000
u
+
= y
+
or
u
+
= 2.44 ln y
+
+ 5.0
0.8 0.4 0.2 0
Webster et al. (1996)
u
+
Figure 22. Time-averaged mean streamwise velocity: , x = 0.5; , x = 1.0;
, x = 1.5; - , x = 2.0; , x = 2.5; ., x = 3.0;
e
, x = 3.5 (exit).
138 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(e) x = 2.5
3
2
1
0
1
0 0.4 0.8 1.2 1.6
y/
R
e
y
n
o
l
d
s

s
t
r
e
s
s
e
s
( f ) x = 3.5 (exit)
3
2
1
0
1
0 0.4 0.8 1.2 1.6
y/
(c) x =1.5
3
2
1
0
1
0 0.4 0.8 1.2 1.6
R
e
y
n
o
l
d
s

s
t
r
e
s
s
e
s
(d) x = 2.0
3
2
1
0
1
0 0.4 0.8 1.2 1.6
(a) x = 0.5
3
2
1
0
1
0 0.4 0.8 1.2 1.6
R
e
y
n
o
l
d
s

s
t
r
e
s
s
e
s
(b) x =1.0
3
2
1
0
1
0 0.4 0.8 1.2 1.6
Figure 23. Time-averaged Reynolds stresses in outer coordinates; , (u
/
2
)
1/2,+
; , (v
/
2
)
1/2,+
;
, (w
/
2
)
1/2,+
; , (u
/
v
/
)
+
; symbols: at-plate boundary layer of Spalart (1988) at Re

= 1410.
Proles of the time-averaged and normalized (on wall parameters) turbulence
kinetic energy production rate
+
= (u
/
v
/
)
+
u
+
/y
+
at eight streamwise stations
are compared to the DNS of Spalart (1988) and the experimental data of Kim, Kline
& Reynolds (1968) in gure 24. Spalart noticed that his DNS proles of
+
at three
dierent momentum thickness Reynolds numbers are self-similar and also agree very
well with Kim et al. He attributed this to the fact that at such relatively low Reynolds
numbers the decrease of Reynolds shear stress and the increase of mean velocity
gradient cancel each other in the product to a remarkable degree. It is clear from
gure 24 that the present computation faithfully reproduces this feature, as evident
in the self-similarity of the proles at x = 2.5, 3.0 and 3.5 and the agreement with
Spalart (1988) and Kim et al. (1968).
Boundary layer transition induced by passing wakes 139
0.1
0 10 20 30 40
y
+
50 60 70 80
0.2
0.3
0.4

+

c
u
+
/
c
y
+
Figure 24. Time-averaged non-dimensional turbulence kinetic energy production near the wall:

,
Spalart (1988);
e
, Kim, Kline & Reynolds (1968); , x = 0.5; O, x = 0.75; , x = 1.0; ,
x = 1.5; - , x = 2.0; , x = 2.5; ., x = 3.0; , x = 3.5 (exit).
In the transition region the peak of
+
is located away from the wall and there is
also a large over-shoot. This is consistent with the turbulence intensity and Reynolds
shear stress proles presented in gure 23. As the ow approaches the end of transition,
the peak shifts towards the wall and the maximum value of
+
drops. Again, note
that the proles are normalized by the local friction velocity. A clearer picture of
the streamwise evolution of the absolute peak values is given in gure 25. This
gure shows local maxima of time-averaged turbulence kinetic energy, wall-normal
uctuations, Reynolds shear stress and turbulence kinetic energy production. A large
over-shoot in turbulence kinetic energy is seen prior to the completion of transition.
The peak is located at x 1.85. The time-averaged wall-normal uctuations do
not show as noticeable an over-shoot as the turbulence kinetic energy. Production
of turbulence kinetic energy and Reynolds shear stress peak at the same location,
though the degree of over-shoot is much stronger for the former. The location where
the Reynolds shear stress attains its maximum value is the same as the time-averaged
mean skin friction, i.e. x 2.15. In the time-averaged sense this also corresponds
to the end of transition as indicated by the results shown in gures 22, 23 and
24.
Overall, the results presented in this section demonstrate that the present simula-
tions yield correct time-averaged ow statistical properties in the laminar regime
near the inlet and the fully turbulent regime near the exit. Between these two
ends time-averaged mean and second-order statistics proles across the boundary
layer seldom vary monotonically with the increase of x. Completion of transition,
in a time-averaged sense, can be dened as where the mean skin friction reaches
its peak value. At this location (Re

660) Reynolds shear stress and produc-


tion of turbulence kinetic energy also attain their peak values. Downstream proles
of the mean and turbulence statistics are in good agreement with previous DNS
and experimental data for fully turbulent at-plate boundary layers. In the work
140 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
0.1
0 1 2
x
3
0.2
0.3
0.4

1
0
2

m
a
x
,

c
u
/
c
y
|
m
a
x

m
a
x
,

m
a
x
(b)
0 1 2 3
0.005
0.010
0.015
(a)
Figure 25. Streamwise evolution of maximum Reynolds stresses: - , turbulence kinetic energy;
, wall-normal uctuation; , shear stress; , production of turbulence kinetic energy.
of Liu & Rodi (1991), the location of transition was considered to be where the
turbulent strips merge, and where the turbulence intensity reaches the level pre-
vailing in turbulent boundary layers. The location of merging is dicult to dene,
since it involves transient behaviour and many other subjective factors. Turbulence
intensities also develop over-shoot characteristics in the process of transition. At
the given passing frequency their estimated transition location is x 1.85. Taking
into account the dierent denitions, there is an overall agreement in the transi-
tion location between the present computation and the experiments of Liu & Rodi
(1991).
Boundary layer transition induced by passing wakes 141
0
0 1 2
x
3
0.002
0.004
0.006

C
f

(b)
0.002
0.5 passing periods
0.7 passing periods
0.9 passing periods
0
0 1 2 3
0.002
0.004
0.006

C
f

(a)
0.002
0 passing periods
0.2 passing periods
0.4 passing periods
Figure 26. Streamwise distributions of phase-averaged skin-friction coecient and free-stream
turbulence kinetic energy; upper curves: (C
f
); lower curves: (K) at y = 0.1.
6. Phase-averaged boundary layer properties
Time averaging is not equivalent to ensemble averaging in this non-stationary ow.
The phase average gives a fuller statistical picture. At t
n

= 0.9 and 0 the wake


is completely outside the (computational) inow plane. At these two instants the
velocity proles simply follow the Blasius solution. Beginning from t
n

= 0 the wake
starts to enter from the top boundary of the inow plane and descends towards
the at plate. Touch-down occurs slightly before t
n

= 0.7. Streamwise variations of


the phase-averaged skin friction and turbulence kinetic energy in the free stream (at
y = 0.1) are shown in gure 26. As the wake passes along the plate its strength is
attenuated, which is also accompanied by a spread in wake width. This eect is seen
in the K versus x plots. The (C
f
) versus x plot shows a local rise in the laminar region
142 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
0.1
0 1 2
x
3
0.2
0.3
0.4

c
f
,

s
t
Figure 27. Streamwise proles of the parameter indicating boundary layer transitional state: ,
based on r.m.s. periodic skin-friction uctuation; , based on r.m.s. periodic Stanton number
uctuation.
when the wake buets the layer. This is distinct from the larger, sustained rise further
downstream where transition occurs, and the boundary layer becomes fully turbulent.
The location of the rise lags the free-stream wake passage as previously discussed.
After the completion of transition no distinct peaks can be seen in the phase-averaged
r.m.s. skin-friction uctuations even though the turbulent layer is still bueted by the
wake. Over the distance where the rise is observed there is a change in the nature
of the uctuations that occurs at around 0.4 passing periods, when the wake reaches
x 0.7. At that point, turbulent spots begin to appear and the uctuation proles
peak deeper in the boundary layer. The appearance changes from that of a bueted
layer to one with self-sustained turbulence. Similar features have also been reported
by Halstead et al. (1997) in multi-stage compressor/turbine measurements.
Figure 27 shows the streamwise distribution of the r.m.s. periodic skin-friction
uctuation coecient
C
f
dened as

C
f
=
__
(C
f
) C
f
_
2
_
1/2
/C
f
. (8)
This coecient is a good indicator for the boundary layers transitional state. Also
presented in the gure is the streamwise distribution of the r.m.s. periodic Stanton
number uctuation coecient
St
, obtained from a heat transfer simulation in which
the wall was slightly heated (Wu & Durbin 1999a). In a laminar ow
C
f
= 0;
after completion of transition in a fully turbulent ow
C
f
0 if the sample size
is suciently large. The denition of
C
f
is precise and does not involve arbitrarily
chosen threshold values. The spike at the inlet station is due to impact of the passing
wake on the at plate. The small peak quickly decays and reaches a local minimum
at x 0.3. The transitional state of the boundary layer at this location is the lowest
prior to the completion of transition. This is again consistent with the notion that
turbulence directly carried by the wake into the near-wall region decays rapidly.
C
f
reaches its global maximum at x 1.2, indicating that at this station the boundary
Boundary layer transition induced by passing wakes 143
(e)
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8
t
n

1
0
2

0.1
0.2
0.3
0.4
1.0
( f )
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8
t
n

0.1
0.2
0.3
0.4
1.0
(c)
1.6
1.2
0.8
0 0.2 0.4 0.6 0.8
1
0
2

0.4
1.0
(d)
1.6
1.2
0.8
0 0.2 0.4 0.6 0.8
0.4
1.0
(a)
0.12
0.08
0.04
0
0 0.2 0.4 0.6 0.8

u
0.04
0.08
0.12
1.0
(b)
0.12
0.08
0.04
0
0 0.2 0.4 0.6 0.8
0.04
0.08
0.12
1.0
Figure 28. Periodic uctuation of streamwise velocity, phase-averaged turbulence kinetic energy
and Reynolds shear stress:
e
, y = 0.002; , y = 0.02; ., y = 0.04; +, y = 0.1; (a, c, e) x = 1.0;
(b, d, f) x = 1.5.
layer has the strongest transitional state, in the sense that the properties of the
boundary layer are most distant from those in either the laminar or turbulent regime.
Figure 27 also shows that within the transition region the degree of upstream and
downstream asymmetry of
C
f
with respect to its peak location is relatively small.
In the remainder of this section we focus on spacetime characteristics of the phase-
averaged mean velocity and second-order turbulence statistics in the transitional
region. Figure 28(a, b) presents the periodic streamwise velocity uctuation (u) u at
x = 1.0 and 1.5 as a function of phase t
n

. At each streamwise station four proles


are shown which are taken from y = 0.002, 0.02, 0.04 and 0.1. Figures 28(c, d) and
gure 28(e, f) present proles of the phase-averaged turbulence kinetic energy (K)
and turbulent shear stress (u
/
v
/
) at the same locations, respectively.
144 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
The highest wall-normal location y = 0.1 in the gure remains outside the time-
averaged boundary layer at both streamwise stations. The variations obtained at
y = 0.1 then simply reect, for the most part, the free-stream wake passing. Proles of
(u) u at the most inner wall-normal location y = 0.002 display sine wave behaviour.
The amplitude of the sine wave motion is strongest when the ow has the largest

C
f
, and decreases towards zero in laminar and fully turbulent regions. The peak in
the sine wave has a phase lag behind the free-stream wake. The variation of (K) at
y = 0.002 shows the statistically averaged eect of turbulent spots. Since at x = 1.0
transition is still in the early stage, (K) and (u
/
v
/
) in gure 28(c, e) return to zero after
the statistically averaged eect of turbulent spots has passed. This is dierent from
the situations shown in gure 28(d, f) where turbulence uctuations do not return
to zero for the whole period because transition is already at its late stage. In the
fully turbulent region (u) u, (K) and (u
/
v
/
) inside the boundary layer approach
their time-averaged mean with little sign of phase dependence. Note that the shear
stress inside the free-stream wake is positive. This is a consequence of the wake
orientation.
In the transitional region, the behaviour of (u) u in the central part of the
boundary layer obtained from y = 0.02 is interesting. Figure 28(b) shows the prole
has two dips within one passing period. These two dips correspond to the two peaks
of (K) at the same wall-normal location shown in gure 28(d). The peak of (K)
between t
n

0.5 and 0.6 is clearly from the free-stream passing wake, and the other
one between t
n

0.9 and 1.0 is from the statistically averaged eect of turbulent


spots. Thus the early dip in (u)u is from the wake decit, and the late dip is because
of the temporal thickening of boundary layer arising from the statistically averaged
eect of turbulent spots. These dips do not appear before the onset, or after the end,
of transition. Liu & Rodi (1991) also discussed similar dips in their experimental
results. They oered two possible explanations, the rst being that upward cylinder
motion at the far end of the rotating disk caused secondary wakes and the second
being that near-wall turbulent spots cause temporal thickening of boundary layer.
Our results demonstrate that the latter is the major cause.
7. Additional simulations and thought experiments
7.1. Eect of periodic velocity uctuation on transition
The experimental measurements of Orth (1993) suggest that periodic velocity uctu-
ations caused by the passing wake have negligible eect on transition. Our thought
experiment presented in this subsection supports his conclusion.
Additional DNS simulations were performed on two hypothetical ows. In the
rst, the wake turbulence was discarded and only the mean proles were fed into
the ow domain. No turbulent uctuations occurred over the computational region.
A phase-averaged mean ow perturbation was seen immediately beneath the wake,
but this was localized and did not lead to instability (which could only be induced
by numerics). A comparison between the phase-averaged velocity distortions under
the zero and normal turbulence wakes is made in gure 29. The boundary layer
lies below y = 0.0076. A weak inection of the instantaneous boundary layer prole
was noticed in the zero intensity case. This transient inection was insucient to
produce breakdown to turbulence inside the boundary layer (in the case when low
level turbulence was added to the wake).
In the second additional simulation the wake turbulence was reduced to 1% of
Boundary layer transition induced by passing wakes 145
0.02
0
0.08 0.04
y
0
0.04
0.06
0.08
0.04 0.08
0.10
u u
Figure 29. Periodic velocity uctuation proles at x = 0.5 in the baseline simulation with full wake
turbulence and in the additional simulation where the inlet wake turbulence is reduced to zero. Full
wake turbulence ( = 0.0086): , t
n

= 0.9; , t
n

= 0.05. Zero wake turbulence ( = 0.0076):


, t
n

= 0.9; - , t
n

= 0.05.
its natural level. Figure 30(a) presents contours of instantaneous streamwise velocity,
u U
ref
over an (x, y)-plane at t/ = 8.0. The velocity contours show no sign
of turbulent spots or of transition. Figure 30(b) presents contours of u
/
over the
(x, z)-plane at y/L = 7.38 10
4
at the same instant. The near-wall disturbances
generated at the inlet evolve toward long-streamwise-wavelength waves, with spanwise
modulation. The wake passage is seen more clearly in the v
/
proles of gure 30(c).
These gures show that the boundary layer lters out short-wave inlet disturbances
(Jacobs & Durbin 1998). They decay rapidly and are not the origin of the small-scale
spots that occur beneath higher free-stream wake turbulence.
7.2. Eect of streamwise pressure gradient on transition
Streamwise pressure gradient exerts an overwhelming inuence on the stability of
a laminar boundary layer. A decrease in pressure in the downstream direction, i.e.
favourable gradient, has a stabilizing eect. In the case of a at plate this causes
the critical Reynolds number to become larger than Re

= 200. An adverse gradient


leads to instability.
The manner in which the wakes are introduced at the inlet causes a localized
streamwise pressure gradient. The inlet velocity distribution is a superposition of
Blasius and wake velocity proles. This distribution is forced to satisfy the no-slip
boundary condition at the inow station by an articial damping through the use
of (7).
It has been suggested in the literature that passing wakes produce streamwise
pressure gradients of opposite sign depending upon the sign of U
cyl
(see gure 2).
Hodson (1985) and Hodson & Dawes (1998) discussed a negative jet concept related
to this subject. They argued that a negative vertical velocity component inside the
146 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
(c)
0.20
0.15
0 0.5
x
z
1.0 3.0
0.10
0.05
1.5 2.0 2.5 3.5
(b)
0.20
0.15
0 0.5
z
1.0 3.0
0.10
0.05
1.5 2.0 2.5 3.5
(a)
0.6
0 0.5
y
1.0 3.0
0.4
0.2
1.5 2.0 2.5 3.5
Figure 30. Instantaneous velocity elds at t/ = 8.0 in the additional simulation where the inlet
wake turbulence is reduced by a factor of 100: (a) u U
ref
over one (x, y)-plane; (b) 10
2
u
/
over the
(x, z)-plane of y = 7.38 10
4
; (c) 10
2
v
/
over the (x, z)-plane of y = 7.38 10
4
.
wake causes a high-pressure stagnation point on the wall. Thus there exists a localized
favourable pressure gradient near the wake. When U
cyl
is positive instead of negative,
they argued that the positive vertical velocity component inside the wake results in
a low-pressure point on the wall, causing a localized adverse pressure gradient near
the wake. However, Hunt, Durbin & Wu (1998) have shown that the pressure change
associated with this negative jet eect is quite small since its magnitude is dependent
to the square of the wake velocity decit. Nevertheless, the reasoning of Hodson
(1985) may still be helpful in understanding the inlet pressure gradient encountered
in the current study.
At the inlet, when U
cyl
< 0 the mass contained inside the negative jet is forced to
enter the computational domain as the wall is approached because the wake decit
is damped by Blasius to satisfy no slip. This results in a localized high pressure.
When U
cyl
> 0 the mass contained inside the negative jet is sucked away from
the computational domain as the wall is approached, thus producing a localized
low pressure at the wall. We will present three additional numerical experiments in
this subsection to demonstrate and quantify the eect of inlet streamwise pressure
gradient on wake-induced boundary layer transition.
Pictures of the three boundary layers, drawn using contours of the instantaneous
streamwise velocity u over one random x, y cross-section, are shown in gure 31.
Boundary layer transition induced by passing wakes 147
(c)
0.8
0.6
0 0.5
x
y
1.0 3.0
0.4
0.2
1.5 2.0 2.5 3.5
(b)
0.8
0.6
0 0.5
y
1.0 3.0
0.4
0.2
1.5 2.0 2.5 3.5
(a)
0.8
0.6
0 0.5
y
1.0 3.0
0.4
0.2
1.5 2.0 2.5 3.5
Figure 31. Instantaneous velocity u over one (x, y)-plane: (a) additional simulation with
U
cyl
= +0.7U
ref
; (b) additional thought experiment with inlet mean wake decit set to zero
(U
cyl
= 0.7); (c) additional thought experiment with inlet mean wake decit set to zero
(U
cyl
= +0.7).
Figure 31(a) corresponds to a simulation in which U
cyl
= +0.7U
ref
as opposed
to 0.7U
ref
in the baseline case. Figure 31(b) and gure 31(c) correspond to two
thought experiments with inlet mean wake decit being set to zero and U
cyl
= _0.7,
respectively. The inlet wakes in gures 31(b) and 31(c) are therefore of shear-free
type, i.e. mean wake velocities (u
wake
)
z
and (v
wake
)
z
have been subtracted from the
corresponding inlet wake proles while retaining only the turbulence uctuations.
Figure 32 shows the time-averaged streamwise pressure coecient in the three
cases, together with that in the baseline case. C
pw
is dened as 2(p
wall
p
wall, inlet
)/U
2
ref
.
It is seen from gure 32 that when inlet mean wake decit is retained, damping using
Blasius of the downward-moving wake results in a mild favourable pressure gradient
in the region x < 0.75. Damping using Blasius of the upward-moving wake produces
a mild adverse pressure gradient in the same region. Also evident from gure 32 is
that a nearly zero streamwise pressure gradient is achieved beyond x = 0.75 for all the
cases, except for near the exit where a slight favourable pressure gradient is visible.
Streamwise pressure gradients in both of the shear-free wake cases are nearly identical
and essentially zero across the whole computational domain. Without a mean wake
148 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
0
0 1
x
C
pw
0.1
0.2
2 3
0.1
0.2
Figure 32. Time-averaged wall static pressure coecient: , baseline case U
cyl
= 0.7; ,
additional simulation U
cyl
= +0.7; , additional simulation with inlet mean wake decit set to
zero (U
cyl
= 0.7); , additional simulation with inlet mean wake decit set to zero (U
cyl
= +0.7).
decit, the no-slip condition forced at the inlet by Blasius damping results in minimal
distortion of the ow.
Figure 33 compares visualizations of the boundary layer transition processes. Shown
in the gure are contours of instantaneous wall-normal velocity uctuation v
/
over the
(x, z)-plane of y = 7.38 10
4
. Fluctuations with magnitude smaller than 0.005 are
grouped into the background. Figure 33(a) shows that under a mild localized adverse
pressure gradient, wake turbulence introduced from the inlet does not exhibit rapid
decay in the near-wall region. A quasi-two-dimensional turbulent strip forms right
from the inlet. No isolated turbulent spots emerge in this instance. At the instant
shown on gure 33(a), three turbulent regions co-exist and are separated streamwise
by relatively quiet laminar-like strips. In the baseline case only two streamwise
separated turbulent regions co-exist at any time despite the fact that the passing
frequency is the same in both cases.
Figures 33(b) and 33(c) visualize the boundary layer transition processes for the
two shear-free wake cases. The emergence of turbulent spots occurs closer to the
inlet than in the baseline case. For both of the shear-free wake cases there are two
full turbulent strips plus isolated turbulent spots over the entire (x, z)-plane. Thus
elimination of the streamwise pressure gradients makes the transition processes nearly
the same, regardless of the wake orientation. Such a characteristic is also manifested
in statistically averaged boundary layer properties. Figure 34 shows the time-averaged
skin-friction coecient for all the four cases. A localized favourable pressure gradient
delays the minimal skin-friction location but with a higher over-shoot. A localized
adverse pressure gradient makes the minimal skin friction occur closer to the inlet but
with a lower C
f
in the turbulent region. Removing the mean wake decit produces
an intermediate skin-friction distribution.
Boundary layer transition induced by passing wakes 149
(c)
0.20
0.15
0 0.5
x
z
1.0 3.0
0.10
0.05
1.5 2.0 2.5 3.5
(b)
0.20
0.15
0 0.5
z
1.0 3.0
0.10
0.05
1.5 2.0 2.5 3.5
(a)
0.20
0.15
0 0.5
z
1.0 3.0
0.10
0.05
1.5 2.0 2.5 3.5
Figure 33. Instantaneous wall-normal uctuation v
/
over the (x, z)-plane of y = 7.38 10
4
: (a)
additional simulation with U
cyl
= +0.7U
ref
; (b) additional thought experiment with inlet mean wake
decit set to zero (U
cyl
= 0.7); (c) additional thought experiment with inlet mean wake decit set
to zero (U
cyl
= +0.7).
8. Summary
A notable feature of the present study is its dual relevance to the fundamental
subject of boundary layer instability under moderate levels of disturbance, and the
engineering subject of turbomachinery aerodynamics.
The concept of pus has proven to be relevant in passing wake-induced bypass tran-
sition. Such structures evolve from near-wall leading-edge (inlet) disturbances through
a receptivity phase. They have a tendency to elongate and decay. Amplication is
observed when certain types of free-stream vortices interact with the boundary layer
ow through a local KelvinHelmholtz-like instability. Breakdown in wake-induced
bypass transition usually occurs in the outer part of the boundary layer, following the
typical KelvinHelmholtz-type wavy motion in the velocity eld. When this happens,
negative streamwise uctuations associated with the inectional proles evolve into
strong forward eddying motions, producing young turbulent spots. Violent upward
motion exists near the interface of the positive/negative streamwise uctuations. This
interface also forms the downstream edge of the turbulent spot. Thought experiments
show mean wake distortion of the boundary layer is less important in receptivity
and transition compared to the interaction between boundary layer and free-stream
150 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
0.002
0 1
x
C
f
2 3
0.004
0.008
0.006
Figure 34. Time-averaged skin-friction coecient:
e
, baseline case U
cyl
= 0.7; ., additional sim-
ulation U
cyl
= +0.7; , additional simulation with inlet mean wake decit set to zero (U
cyl
= 0.7);
O, additional simulation with inlet mean wake decit set to zero (U
cyl
= +0.7); , Blasius solution
without wake; every other 20 points are shown.
turbulent eddies carried by the passing wakes. These ndings have answered, to a
satisfactory degree, all the questions we have raised in , 1.
Visualizations obtained from this study have been compared to the liquid crystal
experiments of Zhong et al. (1998) and Kittichaikaran et al. (1999). Specically,
geometrical characteristics of the simulated and measured pus prior to breakdown,
as well as the matured turbulent spot, are found to be in good agreement. The
turbulent spot has an arrowhead pointing upstream, in the reverse direction to many
previous fundamental studies on boundary layer instability. We discovered that when
breakdown occurs in the outer layer, where the local convection speed is large, as
in the present case, the arrowhead points upstream. This often happens when the
boundary layer is subjected to perturbation by free-stream eddies. When breakdown
occurs near the wall, e.g. as a result of a disturbance articially introduced at the
wall, the turbulent spot has an arrowhead pointing downstream.
We have also studied the wake-induced transition problem from a statistical point
of view. The computed boundary layer properties upstream of the onset of transition
follow the Blasius solution, and those after the completion of transition follow
fully developed canonical turbulent boundary layers. During transition the peak
of turbulence kinetic energy is displaced away from the wall, in agreement with
Alfredsson & Matsubara (1996). Completion of transition is marked by the streamwise
location where the time-averaged mean skin friction reaches its maximum. Reynolds
shear stress and production of turbulence kinetic energy also attain their maxima at
the same location, while turbulence kinetic energy peaks further upstream. During
transition phase-averaged mean streamwise velocity in the near-wall region displays
sinusoidal behaviour in time. The amplitude of such sine wave motion peaks when
the boundary layer is most distant from both laminar and turbulent regimes. In
Boundary layer transition induced by passing wakes 151
the central part of the boundary layer, the phase-averaged mean streamwise velocity
exhibits two dips for each passing cycle, one from the passing wake decit, the other
from the statistically averaged eect of turbulent spot growth. These features are in
agreement with the experiments of Liu & Rodi (1991). Unsteady Reynolds-averaged
NavierStokes (RANS) computations of this wake-induced transitional boundary
layer have also been conducted, and the predictions compared with present DNS
data. That work was reported in Wu & Durbin (1998, 1999a, b).
We thank Charles D. Pierce for the use of his code. This work is supported by the
Academic Strategic Alliance Program of the US Department of Energy Accelerated
Strategic Computing Initiative. The simulations were performed on the Cray T3E at
the Pittsburgh Supercomputing Center (PSC). Discussions with D. S. Henningson,
H. P. Hodson, S. K. Lele, P. Moin, W. C. Reynolds and M. M. Rogers are gratefully
acknowledged. The computer hardware support from M. Fatica and R. Subramanya
(PSC) is gratefully acknowledged. J.C.R.H. is grateful to the Center for Turbulence
Research for support during this work.
REFERENCES
Addison, J. S. & Hodson, H. P. 1990 Unsteady transition in an axial ow turbine, part 1:
measurements on the turbine rotor; part 2: cascade measurements and modeling. Trans. ASME:
J. Turbomachinery 112, 206214.
Akselvoll, K. & Moin, P. 1996 Large eddy simulation of turbulent conned coannular jets. J. Fluid
Mech. 315, 387411.
Alfredsson, P. H. & Matsubara, M. 1996 Streaky structures in transition. In Transitional Boundary
Layers in Aeronautics (ed. R. A. W. M. Henkes & J. L. Ingen), pp. 374386. Elsevier.
Berlin, S., Lundbladh, A. & Henningson, D. S. 1994 Spatial simulations of oblique transition.
Phys. Fluids 6, 19491951.
Coles, D. 1956 The law of the wake in the turbulent boundary layer. J. Fluid Mech. 1, 191226.
Corral, R. & Jimenez, J. 1994 Direct numerical determination of the minimum bypass Reynolds
number in boundary layers. 74th Fluid Dynamics Symp., Chania, Crete, Greece, pp. 19-119-9.
Dong, Y. & Cumpsty, N. A. 1990 Compressor blade boundary layers, part 1: test facility and mea-
surements with no incident wakes; part 2: measurements with incident wakes. Trans. ASME:
J. Turbomachinery 112, 222240.
Dring, R. P., Joslyn, H. D., Hardin, L. W. & Wagner, J. H. 1982 Turbine rotor-stator interaction.
Trans. ASME: J. Engng Power 104, 729742.
Ghosal, S. & Rogers, M. M. 1997 A numerical study of self-similarity in a turbulent plane wake
using large-eddy simulation. Phys. Fluids 9, 17291739.
Goldstein, M. E. & Wundrow, D. W. 1998 On the environmental realizability of algebraically
growing disturbances and their relation to Klebano modes. Theor. Comput. Fluid Dyn. 10,
171186.
Grek, H. R., Kozlov, V. V. & Ramazanov, M. P. 1985 Three types of disturbances from the
point source in the boundary layer. In Laminar Turbulent Transition 2 (ed. V. V. Kozlov), pp.
267272. Springer.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P. & Shin, H. W. 1997
Boundary layer development in axial compressors and turbines part 1: composite picture;
part 2: compressors; part 3: LP turbines; part 4: computations and analysis. Trans. ASME:
J. Turbomachinery 119, 114127, 128138, 225237, 426443.
Hamilton, J. M., Kim, J. & Waleffe, F. 1995 Regeneration mechanisms of near-wall turbulence
structures. J. Fluid Mech. 287, 317348.
Hancock, P. E. & Bradshaw, P. 1989 Turbulence structure of a boundary layer beneath a turbulent
free stream. J. Fluid Mech. 205, 4576.
Henningson, D. S. & Kim, J. 1991 On turbulent spots in plane Poiseuille ow. J. Fluid Mech. 228,
183205.
152 X. Wu, R. G. Jacobs, J. C. R. Hunt and P. A. Durbin
Henningson, D. S., Lundbladh, A. & Johansson, A. V. 1993 A mechanism for bypass transition
from localized disturbances in wall-bounded shear ows. J. Fluid Mech. 250, 169207.
Henningson, D. S., Spalart, P. R. & Kim, J. 1987 Numerical simulations of turbulent spots in
plane Poiseuille and boundary layer ow. Phys. Fluids 30, 29142917.
Hodson, H. P. 1985 Measurements of wake-generated unsteadiness in the rotor passage of axial
ow turbines. Trans. ASME: J. Engng Gas Turbine Power 107, 467476.
Hodson, H. P. 1998 Blade row interactions in low pressure turbines. In Blade Row Interference
Eects in Axial Turbomachinery Stages (ed. C. H. Sieverding & H. P. Hodson). Von Karman
Institute for Fluid Dynamics Lecture Series 1998-02.
Hodson, H. P. & Dawes, W. N. 1998 On the interpretation of measured prole losses in unsteady
wake-turbine blade interaction studies. Trans. ASME: J. Turbomachinery 120, 276284.
Hunt, J. C. R. & Durbin, P. A. 1999 Perturbed vortical layers and shear sheltering. Fluid Dyn. Res.
24, 375404.
Hunt, J. C. R., Durbin, P. A. & Wu, X. 1998 Interaction between free-stream turbulence and
boundary layers. In Annual Research Briefs, pp. 113124. Center for Turbulence Research,
Stanford University,
Jacobs, R. G. & Durbin, P. A. 1998 Shear sheltering and the continuous spectrum of the Orr
Sommerfeld equation. Phys. Fluids 10, 20062011.
Jahanmiri, M., Prabhu, A. & Narasimha, R. 1996 Experimental studies of a distorted turbulent
spot in a three-dimensional ow. J. Fluid Mech. 329, 124.
Kim, H. T., Kline, S. J. & Reynolds, W. C. 1968 An experimental study of turbulence production
near a smooth wall in a turbulent boundary layer with zero pressure gradient. Rep. MD-20.
Stanford University.
Kittichaikarn, C., Ireland, P. T., Zhong, S. & Hodson, H. P. 1999 An investigation on the onset
of wake-induced transition and turbulent spot production rate using thermochromic liquid
crystals. ASME Turbo Expo 99, Indianapolis, Indiana (submitted).
Klingmann, B. G. B. 1992 On transition due to three-dimensional disturbances in plane Poiseuille
ow. J. Fluid Mech. 240, 167195.
Korakianitis, T. 1993 On the propagation of viscous wakes and potential ow in axial-turbine
cascades. Trans. ASME: J. Turbomachinery 115, 118127.
Leib, S. J., Wundrow, D. W. & Goldstein, M. E. 1999 Eect of free-stream turbulence and other
vortical disturbances on a laminar boundary layer. J. Fluid Mech. 380, 169203.
Liu, X. & Rodi, W. 1991 Experiments on transitional boundary layers with wake-induced unsteadi-
ness. J. Fluid Mech. 231, 229256.
Mayle, R. E. 1991 The role of laminar turbulent transition in gas turbine engines. Trans. ASME:
J. Turbomachinery 113, 509537.
Mayle, R. E. & Dullenkopf, K. 1991 More on the turbulent-strip theory for wake-induced
transition. Trans. ASME: J. Turbomachinery 113, 428432.
Moser, R. D., Rogers, M. M. & Ewing, D. W. 1998 Direct simulation of a self-similar plane wake.
J. Fluid Mech. 368, 255289.
Orth, U. 1993 Unsteady boundary layer transition in ow periodically disturbed by wakes.
Trans. ASME: J. Turbomachinery 115, 707713.
Pfeil, H., Herbst, R. & Schroder, T. 1983 Investigation of the laminar turbulent transition of
boundary layers disturbed by wakes. Trans. ASME: J. Engng Power 105, 130137.
Pierce, C. D. & Moin, P. 1998 Large eddy simulation of a conned coaxial jet with swirl and heat
release. AIAA Paper 98-2892.
Rai, M. M. & Moin, P. 1993 Direct numerical simulation of transition and turbulence in a spatially
evolving boundary layer. J. Comput. Phys. 109, 169192.
Raj, R. & Lakshminarayna, B. 1973 Characteristics of the wake behind a cascade of airfoils.
J. Fluid Mech. 61, 707730.
Schlichting, H. 1979 Boundary Layer Theory, 7th Edn. McGraw-Hill.
Spalart, P. R. 1988 Direct simulation of a turbulent boundary layer up to Re

= 1410. J. Fluid
Mech. 187, 6198.
Spalart, P. R., Moser, R. D. & Rogers, M. M. 1991 Spectral methods for the NavierStokes
equations with one innite and two periodic directions. J. Comput. Phys. 96, 297324.
Walker, G. J. 1993 The role of laminar turbulent transition in gas turbine engines: a discussion.
Trans. ASME: J. Turbomachinery 115, 207217.
Boundary layer transition induced by passing wakes 153
Webster, D. R., DeGraaff, D. B. & Eaton, J. K. 1996 Turbulence characteristics of a boundary
layer over a two-dimensional bump. J. Fluid Mech. 320, 5369.
Westin, K. J. A., Boiko, A. V., Klingmann, B. G. B., Kozlov, V. V. & Alfredsson, P. H. 1994
Experiments in a boundary layer subjected to free-stream turbulence. Part 1. Boundary layer
structure and receptivity. J. Fluid Mech. 281, 193218.
Wu, X. & Durbin, P. A. 1998 Boundary layer transition induced by periodic wakes. Trans.
ASME: J. Turbomachinery. (submitted).
Wu, X. & Durbin, P. A. 1999a Numerical simulation of heat transfer in a transitional boundary
layer with passing wakes. Trans. ASME: J. Heat Transfer (submitted).
Wu, X. & Durbin, P. A. 1999b Numerical experiments and modeling of the transitional and
turbulent boundary layers induced by periodically passing wakes. Rep. Center for Integrated
Turbulence Simulation, Stanford University.
Wu, X. & Squires, K. D. 1997 Large eddy simulation of an equilibrium three-dimensional turbulent
boundary layer. AIAA J. 35, 6774.
Yang, K. S., Spalart, P. R. & Ferziger, J. H. 1992 Numerical studies of natural transition in a
decelerating boundary layer. J. Fluid Mech. 240, 433468.
Zhong, S., Kittichaikarn, C., Hodson, H. P. & Ireland, P. T. 1998 Visualization of turbulent spots
under the inuence of adverse pressure gradients. In Proc. 8th Intl Conf. on Flow Visualization,
Italy.

You might also like